phenyltrichlorosilane: Self-Polymerization within Spatially Confined

May 3, 2013 - Our Lady of the Lake College, 5414 Brittany Drive, Baton Rouge, Louisiana 70808, United States. •S Supporting Information. ABSTRACT: T...
0 downloads 0 Views 982KB Size
Article pubs.acs.org/Langmuir

Directed Surface Assembly of 4‑(Chloromethyl)phenyltrichlorosilane: Self-Polymerization within Spatially Confined Sites of Si(111) Viewed by Atomic Force Microscopy Tian Tian,† Zorabel M. LeJeune,‡ and Jayne C. Garno*,† †

Department of Chemistry, Louisiana State University, 232 Choppin Hall, Baton Rouge, Louisiana 70803, United States Our Lady of the Lake College, 5414 Brittany Drive, Baton Rouge, Louisiana 70808, United States



S Supporting Information *

ABSTRACT: The self-polymerization of 4-chloromethylphenyltrichlorosilane (CMPS) was studied within spatially confined nanoholes on Si(111) using atomic force microscopy (AFM). Surface platforms of nanoholes were fabricated within a film of octadecyltrichlorosilane using immersion particle lithography. A heating step was developed to temporarily solder the silica mesospheres to the surface, to enable sustained immersion of mesoparticle masks in solvent solutions for the particle lithography protocol. Substrates with a film of mesospheres were heated briefly to anneal the particles to the surface, followed by a rinsing step with sonication to remove the silica beads to generate nanopores within an octadecyltrichlorosilane (OTS) film. Nanopatterned surface templates were immersed in CMPS solutions and removed at different time points to monitor the successive growth of nanostructures over time. Analysis of AFM images after progressive exposure of the nanoholes to solutions of CMPS provided quantitative information and details of the surface self-assembly reaction. Pillar nanostructures of CMPS with different heights and diameters were produced exclusively within the exposed areas of the substrates. Throughout the reaction, the surrounding matrix of OTS-passivated substrate did not evidence growth of CMPS; the surface assembly of CMPS was strictly confined within the nanopores. The diameter of the CMPS nanostructures grew to match the initial sizes of the confined areas of Si(111) but did not spread out beyond the edges of the OTS nanocontainers. However, the vertical growth of columns was affected by the initial size of the sites of uncovered substrate, evidencing a direct correspondence; larger sites produced taller structures, and correspondingly the growth of shorter structures was observed within smaller nanoholes. The heights of CMPS nanostructures indicate that multilayers were formed, with taller columns generated after longer immersion times. These experiments offer intriguing possibilities for using particle lithography as a general approach for nanoscale studies of molecular self-assembly.



beam lithography,12 and X-ray lithography.13 The benzyl halide surfaces of 4-chloromethylphenyltrichlorosilane (CMPS) furnish sites for nucleophilic substitution reactions,14 provide ligands for binding DNA,9 polymers,15 peptide synthesis,16 fluorescent binding assays,17 and chromophores.18 There are only a few methods of positioning molecules at a local scale of nanometers that will enable studies with atomic force microscopy (AFM) at the molecular level. Methods of scanning probe-based lithography that have been used to create patterns of organosilanes include bias-induced lithography,19 nanoshaving,20,21 nanografting,22 Dip-Pen nanolithography,23,24 and constructive nanolithography.25,26 Although the size, shape, and terminal group of the patterns can be exquisitely controlled, a limitation of scanning probe-based approaches is that each pattern is written or inscribed individually by a relatively slow, serial process. To scale up to produce millions of nanopatterns with high density, we have been working to develop and refine methods using particle lithography. Particle

INTRODUCTION Model systems of n-alkanethiols have been well-studied, including the surface self-assembly mechanisms, structures, and growth.1−3 Organosilane self-assembled monolayers (SAMs) were first introduced by Sagiv in 1980,4 which offer the advantage of not requiring substrates comprised of expensive precious metals. In particular, organosilane SAMs can be formed on glass and transparent surfaces for sensing applications. The exact mechanism of the surface assembly of organosilanes is more complicated than for n-alkanethiols and remains a target for investigation.5 Organosilanes attach to oxidized surfaces mediated by steps of hydrolysis and condensation to form cross-linked films. Competitive reactions with adjacent molecules are difficult to control, often generating multilayer films. From an applications perspective, generating interfaces of well-defined structure and composition is critical for emerging technologies based on molecularly thin organic films. Aromatic organosilanes have previously been studied as surface layers for lithography protocols such as deep UV photo irradiation,6−9 near field scanning optical microscopy,9 microcontact printing,10 scanning tunneling microscopy,11 electron © XXXX American Chemical Society

Received: March 18, 2013 Revised: May 1, 2013

A

dx.doi.org/10.1021/la4010032 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

lithography has also been referred to as nanosphere lithography27 or colloidal lithography.28 Particle lithography has been applied successfully to pattern proteins,29−34 metals,35−40 polymers,41,42 nanoparticles,43−45 and other inorganic materials.46,47 For particle lithography, mesospheres are used as a surface mask to control the deposition of molecules and nanomaterials. Innovative protocols with particle lithography have recently been developed to pattern thiol48−50 and organosilane51 SAMs, which enables exquisite control of the chemical functionalities presented at interfaces. The periodicity and density of SAM patterns can be designed by selecting the diameters of mesospheres used for patterning.52 By combining particle lithography with different deposition methods, patterns such as rings, pores, or dot nanostructures can be produced.53 In this Article, a protocol for particle lithography using immersion was applied to study the surface self-assembly of a 4chloromethylphenyltrichlorosilane from solution. Over time, CMPS spontaneously forms multilayered surface structures through hydrolysis of Si−Cl bonds to form trisilanols, which bridge into cross-linked Si−O−Si networks. Designed surface platforms with well-defined sizes of enclosed reaction sites enabled AFM characterizations of surface changes at the nanoscale for samples prepared ex situ. Typically, mesospheres detach upon immersion in liquids. To address this problem, a strategy of annealing the masks of silica mesoparticles was developed to prepare nanocontainers within a film of octadecyltrichlorosilane (OTS). The natural variations in the sizes of the containers provide a snapshot of the reaction progress at defined intervals up to 20 h after CMPS immersion, with fixed conditions of temperature, humidity, and concentration. Controlling the selectivity and dimensions of surface sites for subsequently assembling supramolecular structures will provide information to elucidate mechanisms and kinetics of surface reactions.



Figure 1. General steps for immersion particle lithography. (A) A mask of silica mesospheres was deposited on the surface of Si(111). (B) After the mesospheres were rinsed away, a film of OTS with nanopores was formed on the substrate. (C) The nanopores were backfilled with CMPS by an immersion step.

EXPERIMENTAL SECTION

Atomic Force Microscopy (AFM). Scanning probe microscope models 5500 and 5420 (Agilent Technologies, Chandler, AZ) were used for characterizing samples. The AFM images in Figures 2, 4, 5, and 6 were acquired with tapping mode in air using silicon nitride tips with a spring constant of 48 N/m and average resonant frequency of 176 kHz (Nanoscience Instruments, Phoenix, AZ). Figure 3 was obtained using contact mode imaging with silicon nitride tips with an average spring constant of 0.5 N/m (MSCT, Veeco Metrology, Santa Barbara, CA). Images were processed using Gwyddion.54 Nanoshaving. For nanoshaving, a higher force was applied to the AFM tip (ranging from 2 to 10 nN) to push the probe through the matrix film to make contact with the substrate. A nanoshaved pattern was generated by scanning over a small area several times, while applying a force higher than that used for imaging. The local pressure at the area of contact produced sufficient shearing force to displace adsorbates during scanning. The area was swept 10 times in a raster pattern in air. Afterward, the same AFM probe was used to characterize the nanoshaved areas in situ by returning to a lower force setting. Immersion Particle Lithography. The general procedure for particle lithography is outlined in Figure 1. Silicon wafers (Virginia Semiconductor, Frederickburg, VA) were cut into small pieces (1 × 1 cm2) for use as substrates. Surfaces were cleaned by immersion in piranha solution for 30 min. Piranha is a mixture of sulfuric acid and hydrogen peroxide at a (v/v) ratio of 3:1. Caution: Piranha solution is highly corrosive and should be handled carefully. First, a drop of monodisperse silica mesospheres was deposited on Si(111) and dried (Figure 1A). To enable sustained immersion in solvent solutions with mesoparticle masks, a heating step was developed to solder the beads

to the substrate. The samples were heated briefly to anneal the mesospheres to the surface (100 °C for 15 min), before immersion into OTS solutions. The annealed films of mesospheres were used as masks for surface patterning. The samples were cooled to room temperature (25 °C), then immersed into 0.1% octadecyltrichlorosilane (Gelest, Morrisville, PA) in anhydrous toluene for 12 h. Silane molecules assembled on the substrate except in the areas where mesospheres were attached to the surface. Next, the samples were rinsed copiously with deionized water and sonicated with ethanol to remove the silica mesospheres (Figure 1B). The center-to-center spacing between the nanopores corresponds to the diameter of the mesosphere masks. In the final step, the nanopatterned samples were submerged into a 0.006 M solution of 4-chloromethylphenyltrichlorosilane (Gelest, Morrisville, PA) in anhydrous toluene (Figure 1C). Samples were removed at successive intervals to evaluate surface changes over time (30 min, 1 h, 10 h, 20 h). Samples were rinsed with acetone and chloroform with sonication and dried under argon. The uncovered areas of Si(111) that had been masked by mesospheres provided well-defined surface sites for directing the subsequent attachment and growth of CMPS. Samples and experiments were accomplished in ambient conditions (25 °C, ∼50% relative humidity), and humidity and temperature were not specifically controlled.55



RESULTS AND DISCUSSION Studies of molecular self-assembly and intermolecular interactions are critical in the field of supramolecular chemistry. Our experimental strategy combines the local spatial resolution of particle lithography with molecular self-assembly to prepare B

dx.doi.org/10.1021/la4010032 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

Figure 2. Nanopores within a film of OTS viewed by AFM images. (A) Topography frame, 2 × 2 μm2; (B) corresponding phase image; (C) higher magnification topograph, 0.5 × 0.5 μm2; (D) phase image; (E) cursor profile across two patterns traced in (C); and (F) view of a single nanopore.

image reveals that the area of the nanopores measures 1% of the surface. Cursor profiles across two of the nanopores indicate the local thickness of the OTS film ranges from 1.5 to 1.7 nm (Figure 2E). A view of a single nanohole is presented in Figure 2F, revealing the surface texture of the surrounding OTS domains. To further evaluate the thickness of the OTS film, a protocol known as “nanoshaving” was accomplished by applying high mechanical force to the AFM tip to sweep away the OTS film from a selected area. A square pattern was produced by nanoshaving in air as shown in Figure 3A. The pattern

arrays of nanostructures with designated periodicity. Millions of nearly regular nanopatterns can be formed using basic steps of particle lithography (mixing, rinsing, drying, heating, centrifuging, and sonication) to enable exquisite nanoscale control of the geometry, density, and interfacial chemistry of surfaces. For this study, nanopore structures within a film of OTS were used as nanocontainers to designate sites for the growth of CMPS. Of course, at the nanoscale there are small variations in the geometry and sizes of the nanocontainers that are produced. This provides an opportunity to evaluate the size-dependent spatial effects of confinement as a function of exposed surface area. Confined Nanocontainers. Views of the surface structures that were designed to be used as nanocontainers are shown in Figure 2. Nanopores or holes within an OTS thin film were prepared on Si(111) using particle lithography combined with solution immersion. Within an area of 2 × 2 μm2 there are 48 holes, measuring 1.2 ± 0.2 nm in depth (Figure 2A), with an average surface area of 0.003 ± 0.001 μm2. The holes are the sites where the silica mesospheres were displaced from surface. The grooves in the background are due to the natural roughness of polished silicon wafers. The imperfections of the substrate influence the order and periodicity of the mesosphere masks, as well as the uniformity of the pore geometries. The simultaneously acquired phase image (Figure 2B) more clearly defines the shape of the sites of uncovered substrate. A close-up view of three nanostructures (Figure 2C) reveals that the shapes of the nanoholes are not always circular. The center-to-center spacing of the holes measures 250 nm, which matches the diameter of the silica mesospheres used as a patterning mask. The shapes of the nanocontainers are smaller in the phase image (Figure 2D) as compared to the corresponding topography frame because the height images do not distinguish defined edges of the nanopatterns as clearly and measurements include convolution effects of the tip shape. Using the topography images, the surface coverage of uncovered sites measures 2.7%, whereas the lateral force

Figure 3. Square pattern nanoshaved within the OTS film with nanopores. (A) Topography, 2.5 × 2.5 μm2; (B) cursor profile across the nanopores and nanoshaved area.

measures 500 × 500 nm2 and has a depth that is similar to that of the nanopores, ∼1.2 ± 0.2 nm, which is shorter than the value expected for a densely packed SAM of OTS.56 The depth of the pattern and holes are compared side-by-side with the cursor profile in Figure 3B. The thickness of OTS films from ellipsometry measurements has been reported to range from to 2.25 to 2.81 nm for densely packed monolayers formed on silicon substrates.57−59 In a dense arrangement, the alkyl chains of OTS adopt an all-trans C

dx.doi.org/10.1021/la4010032 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

used as a height ruler. A representative cursor measurement across two CMPS nanostructures reveals the heights and lateral dimensions of backfilled CMPS (Figure 4D). The nanostructures are approximately 0.5 ± 0.2 nm taller than the OTS matrix; therefore, the overall thickness would measure 2.1 ± 0.2 nm. Because the theoretical length of CMPS is 0.75 nm,60 the measured height corresponds approximately to a trilayer of CMPS formed after 30 min. After longer immersion, the nanostructures increased in height and width, as shown in Figure 5. A representative

configuration with tilt angles that range from 0° to 17°. The range of measured values can be attributed to changes in surface coverage as well as differences for the methods of sample preparation for OTS. Immersion of a substrate in solvents is the most common approach for preparing films of organosilanes and has produced the most consistent thickness of a monolayer. Backfilling Nanosized Containers with CMPS. By backfilling nanopores produced with particle lithography, exquisitely tiny spatially confined surface sites can be used for studying successive changes after further reaction steps. The combination of chemical synthesis combined with surface engineering likewise provides a unique opportunity for studying spatial confinement effects for surface-based chemical reactions at the molecular level within well-defined nanoscopic areas. Surface patterns of organic thin films can be used as confined nanocontainers for building supramolecular structures through sequential chemical reactions. Successive changes of the surface topography can be viewed after each reaction step. The progressive surface changes during the growth of CMPS nanostructures were characterized with high-resolution AFM at different time intervals to reveal molecular-level details of the surface assembly and self-polymerization of CMPS. The sample with nanopores within an OTS film on Si(111) was immersed in CMPS/toluene solution and removed after 30 min (Figure 4). The nanostructures of CMPS initially form small islands

Figure 5. Surface changes after 1 h immersion in CMPS. (A) AFM topograph, 1 × 1 μm2; (B) view of a single structure, 200 × 200 nm2; (C) topograph, 400 × 400 nm2; and (D) cursor profile across two patterns traced in (C).

topography image (Figure 5A) after 1 h immersion reveals the periodic arrangement of 14 nanostructures of CMPS formed on Si(111) inside the OTS nanocontainers. The heights of the nanostructures are not identical at the nanoscale; smaller nanopores appear to have formed smaller nanostructures. Details are more clearly viewed in Figure 5B for a single nanostructure of CMPS. There is a dark ring surrounding the CMPS nanostructure, indicating that the CMPS did not fully fill the nanoholes and avoided growth at the edges near OTS sidewalls. There are multiple tips at the apex of the nanostructure, resulting from additional nucleation sites being formed during growth. A view of three nanostructures is presented in Figure 5C. The cursor profile (Figure 5D) discloses heights measuring 3 and 8 nm above the OTS matrix, corresponding to 4−10 layers of CMPS. The nanostructures became taller and wider in dimension, according to the initial size of surface sites. To assess whether the self-polymerization of reactive chloro groups of CMPS continued with extended immersion, later time points were evaluated. After 20 h immersion in CMPS, the nanostructures were observed to increase further in length and width as shown in Figure 6. Larger cluster-shaped nanostructures are observed throughout the sample within the OTS nanopores. A representative topograph is presented in Figure 6A showing 14 nanostructures. The growth of CMPS remains confined within the sites of the nanopores, and adsorption is not detected on surrounding matrix areas passivated with OTS.

Figure 4. Nanostructures of CMPS after 30 min immersion. (A) Topography frame, 2.5 × 2.5 μm2; (B) corresponding phase image; (C) zoom-in view of CMPS nanostructures, 0.6 × 0.6 μm2; and (D) cursor profile for the line in (C).

within the central areas of the nanopores and have a distinct boundary surrounding the edges near the inner walls of OTS. Nearly all of the pores evidence growth of CMPS nanostructures (Figure 4A−C). After 30 min, the CMPS has not completely filled the nanopores; however, the height of the CMPS structures corresponds to a multilayer that is taller than the initial height of the surrounding OTS film. The Si(111) substrate can no longer be distinguished to reference as a baseline for height measurements; therefore, the OTS matrix is D

dx.doi.org/10.1021/la4010032 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

Figure 6. Spatially contained nanostructures of CMPS formed after 20 h immersion. (A) Topograph, 1 × 1 μm2; (B) close-up view of a single nanostructure; (C) topograph, 0.5 × 0.5 μm2; and (D) cursor profile across two patterns traced in (C). Figure 7. Conceptual model of the self-assembly of CMPS.

A single CMPS nanostructure is shown in Figure 6B, revealing that multiple nucleation sites were formed over time. The edges of the OTS can be clearly resolved, indicating that CMPS growth is directed in the vertical direction without branching in lateral directions beyond the borders of the nanopores. A view of three nanostructures in Figure 6C reveals that the structures have grown taller and slightly wider, to mostly fill the OTS nanopores. The heights of two of the nanostructures measure 18 and 20 nm above the matrix SAM of OTS (Figure 6D). Using a measured value of 1.5 nm for the thickness of the OTS layer, and an estimated size of 0.75 nm for CMPS, this corresponds to 26−29 molecular layers of CMPS. Analysis of Size Changes for CMPS Nanostructures. Initially, CMPS molecules started to grow at nucleation sites near the center of the nanoholes, and a cross-linking reaction produced multilayers over time in a vertical direction. An approximate model of the self-polymerization scheme is shown in Figure 7, as was previously proposed by Brandow et al.60 The actual molecular assembly is not nearly as organized and wellstructured as pictured in the cartoon; multiple, tangled strands of polymer would be formed. The CMPS nanostructures grow continuously through hydrolysis of the Si−Cl groups to form silanols to produce a cross-linked network. A trace amount of water is needed to convert the chloro groups to hydroxyl groups. Our samples were prepared using anhydrous toluene to minimize the amount of water to that produced by ambient humidity. Measurements of the heights and surface area of CMPS nanostructures after different intervals of immersion are summarized in Table 1. The values for overall thickness of the nanostructures presented in the table include an estimated thickness of 1.5 nm for the thickness of the OTS matrix film. Both the height and the surface area increased as time progressed; however, any polymer branching was constrained by the sides of the nanocontainers. The heights indicate multilayers were formed over time, with taller structures produced by longer immersion. Therefore, CMPS growth was shown to primarily increase in a vertical direction, and larger

Table 1. Surface Changes after Different Intervals of Immersion in CMPS (n = 60) immersion time (h) 0.5 1 10 20

thickness (nm)

thickness range (nm)

± ± ± ±

1.9−5.1 2.8−11 6.3−16 12−33

3.9 7.4 10 20

2.1 2.9 1.4 3.4

average surface area (μm2)

surface area range (μm2)

± ± ± ±

0.0004−0.001 0.001−0.008 0.001−0.01 0.007−0.02

0.0005 0.006 0.008 0.010

0.0002 0.005 0.001 0.005

surface sites produced structures with wider diameters. Histograms of the thickness measurements for the four experiments are provided in the Supporting Information, Figure S1. The correspondence of the nanostructure growth to the initial sizes of nanopores is indicated in Figure 8. The heights of

Figure 8. Correlation between the heights of CMPS nanostructures versus the surface area of OTS nanopores.

CMPS nanostructures versus the surface area of the nanopores were plotted for experiments with immersion times of 1 and 20 h. The trends indicate that larger surface sites produce taller structures, and correspondingly the growth of shorter structures was observed for smaller surface sites. As time progressed, CMPS nanostructures filled the areas within the holes but did not spread out beyond the edges of the OTS nanocontainers. The methyl-terminated headgroups of OTS provided an E

dx.doi.org/10.1021/la4010032 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

the distribution of free silanol sites when CMPS multilayers are formed.60,61 The design for our ex situ experiments primarily evaluated surface changes as a function of immersion time using fixed conditions of concentration, solvent, humidity, and temperature. After designated time intervals, the actual morphology changes were visualized at the nanoscale using ambient AFM characterizations. Future directions would be to prepare even smaller nanocontainers for similar surface studies of CMPS growth, concurrent with studying effects of temperature and solvent parameters, especially considering that we have observed that small changes in solution temperature of only 10 °C can substantially affect the rate of growth for organosilanes. In particular, we plan to study the growth of other organosilanes in nanocontainers, such as molecules with different lengths of alkyl backbones or with different reactive groups. The specific solvents to be investigated will compare organosilane growth in aromatic solvents (e.g., toluene, bicyclohexyl) versus that in alkane-based solvents such as hexane to evaluate if an anisotropic growth mode is possibly mediated through solvent interactions. Another direction for this research would be to use the reactive benzyl halide sites of the CMPS nanostructures to append substituents through nucleophilic substitution reactions to build heterogeneous assemblies.

effective surface mask to prevent nonspecific adsorption of CMPS. It is well-known that chemisorption of trichlorosilanes from solution onto oxide surfaces is influenced by factors such as the nature of the solvent and substrate, temperature, humidity, concentration, and reaction/adsorption time.5,60 For this study, the additional parameter of spatial restriction can be evaluated due to the natural variability of nanoscale patterning by approaches with particle lithography. The diameters of the nanopores ranged from 40 to 100 nm, as measured from the lateral force frames. The diameters of the spheres used to prepare the OTS nanopatterns measured 250 nm; however, the actual size of the area of contact that produced the nanopores is much smaller than the periodicity. Considering that the lateral size of a CMPS molecule is ∼24 ± 2 Å, we calculate that 1600− 80 000 molecules of CMPS can be squeezed within the areas of the different size ranges of the nanopores, estimated for the base layer of the nanostructures. A plot of the approximate number of layers of CMPS that were grown over time is shown in Figure 9. Within the first 60 min, the nanostructures of



CONCLUSIONS An approach based on particle lithography was tested for nanoscale studies of CMPS surface reactions, using test platforms of well-defined nanocontainers within a passivating OTS resist. Details of the surface assembly and subsequent selfpolymerization of CMPS within confined, nanoscopic areas were studied ex situ using AFM. As time progressed, the heights of CMPS nanostructures increased according to the initial sizes of the surface sites of OTS nanoholes. Multilayers were shown to form over time intervals of 0.5−20 h. Further directions for this research will be to study different designs of organosilane molecules and to use different experimental conditions, to gain insight for the dynamics and mechanisms of surface selfassembly reactions.

Figure 9. Estimated numbers of layers of CMPS that formed over time.

CMPS grew at a rate of approximately one layer added every 10 min (using a theoretical value of 0.75 nm/layer for CMPS dimensions). The continuing growth slowed to a rate of about 1−1.3 layers added during each hour after the first 60 min of reaction. Despite multiple experiments to verify the reproducibility of results observed for CMPS growth over time, we did not detect branching and spreading of nanostructures or mushroom shapes. To evaluate the shapes more quantitatively, we compared the surface coverage of the bare areas within OTS before immersion steps (3%) to the surface coverage of the CMPS nanostructures after 20 h of growth (5%). The values are quite comparable, showing little overlap beyond the edges of the nanoholes. The observation of anisotropic growth into columnar shapes that become much taller than the thickness of the surrounding OTS monolayer is an unexpected result. One would expect that the cross-linking reactions between strands during self-polymerization reactions would lead to branching and thus widen the nanostructures to extend in lateral directions beyond the boundaries of the nanoholes. It is possible that the aromatic rings of the solvent mediate a direction for growth through the influence of π−π and coplanar interactions; however, further experiments need to be designed to evaluate the effects of solvent media for CMPS growth. It has been previously postulated that noncovalent interactions between the rings of CMPS and aromatic solvents influence



ASSOCIATED CONTENT

S Supporting Information *

Histograms of the measurements (n = 60) of the thickness of CMPS nanostructures are shown for the individual experiments of Table 1. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*Fax: 225-578-3458. E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS

This research was supported by the National Science Foundation Career program (CHE-0847291, PECASE award) and by the Dreyfus foundation, Camille Dreyfus TeacherScholar award. We also thank the reviewers for helpful insight and suggestions. F

dx.doi.org/10.1021/la4010032 | Langmuir XXXX, XXX, XXX−XXX

Langmuir



Article

(20) Rosa, L. G.; Jiang, J.; Lima, O. V.; Xiao, J.; Utreras, E.; Dowben, P. A.; Tan, L. Selective nanoshaving of self-assembled monolayers of 2(4-pyridylethyl)triethoxysilane. Mater. Lett. 2009, 63, 961−964. (21) Headrick, J. E.; Armstrong, M.; Cratty, J.; Hammond, S.; Sheriff, B. A.; Berrie, C. L. Nanoscale patterning of alkyl monolayers on silicon using the atomic force microscope. Langmuir 2005, 21, 4117−4122. (22) Lee, M. V.; Nelson, K. A.; Hutchins, L.; Becerril, H. A.; Cosby, S. T.; Blood, J. C.; Wheeler, D. R.; Davis, R. C.; Woolley, A. T.; Harb, J. N.; Linford, M. R. Nanografting of silanes on silicon dioxide with applications to DNA localization and copper electroless deposition. Chem. Mater. 2007, 19, 5052−5054. (23) Demers, L. M.; Ginger, D. S.; Park, S. J.; Li, Z.; Chung, S. W.; Mirkin, C. A. Direct patterning of modified oligonucleotides on metals and insulators by dip-pen nanolithography. Science 2002, 296, 1836− 1838. (24) Jung, H.; Kulkarni, R.; Collier, C. P. Dip-pen nanolithography of reactive alkoxysilanes on glass. J. Am. Chem. Soc. 2003, 125, 12096− 12097. (25) Maoz, R.; Frydman, E.; Cohen, S. R.; Sagiv, J. “Constructive nanolithography”: Inert monolayers as patternable templates for in-situ nanofabrication of metal-semiconductor-organic surface structures - A generic approach. Adv. Mater. 2000, 12, 725−731. (26) Wouters, D.; Schubert, U. S. Nanolithography and nanochemistry: Probe-related patterning techniques and chemical modification for nanometer-sized devices. Angew. Chem., Int. Ed. 2004, 43, 2480−2495. (27) Wei, X. Recent developments in the fabrication of ordered nanostructure arrays based on nanosphere lithography. Recent Pat. Nanotechnol. 2010, 4, 194−204. (28) Yang, S. M.; Jang, S. G.; Choi, D. G.; Kim, S.; Yu, H. K. Nanomachining by colloidal lithography. Small 2006, 2, 458−475. (29) Mornet, S.; Bretagnol, F.; Mannelli, I.; Valsesia, A.; Sirghi, L.; Colpo, P.; Rossi, F. Large-scale fabrication of Bifunctional nanostructured polymer surfaces for selective biomolecular adhesion. Small 2008, 4, 1919−1924. (30) Taylor, Z. R.; Patel, K.; Spain, T. G.; Keay, J. C.; Jernigen, J. D.; Sanchez, E. S.; Grady, B. P.; Johnson, M. B.; Schmidtke, D. W. Fabrication of protein dot arrays via particle lithography. Langmuir 2009, 25, 10932−10938. (31) Agheli, H.; Malmstrom, J.; Larsson, E. M.; Textor, M.; Sutherland, D. S. Large area protein nanopatterning for biological applications. Nano Lett. 2006, 6, 1165−1171. (32) Wollenberg, L. A.; Jett, J. E.; Wu, Y.; Flora, D. R.; Wu, N.; Tracy, T. S.; Gannett, P. M. Selective filling of nanowells in nanowell arrays fabricated using polystyrene nanosphere lithography with cytochrome P450 enzymes. Nanotechnology 2012, 23, 385101. (33) Li, J. R.; Henry, G. C.; Garno, J. C. Fabrication of nanopatterned films of bovine serum albumin and staphylococcal protein A using latex particle lithography. Analyst 2006, 131, 244−250. (34) Cai, Y. G.; Ocko, B. M. Large-scale fabrication of protein nanoarrays based on nanosphere lithography. Langmuir 2005, 21, 9274−9279. (35) Haynes, C. L.; van Duyne, R. P. Nanosphere lithography: A versatile nanofabrication tool for studies of size-dependent nanoparticle optics. J. Phys. Chem. B 2001, 105, 5599−5611. (36) Zhou, C. M.; Gall, D. Surface patterning by nanosphere lithography for layer growth with ordered pores. Thin Solid Films 2007, 516, 433−437. (37) Wang, W.; Dai, Z.; Sun, Y.; Sun, Y. Enhancement of optical nonlinearity in binary metal−nanoparticle arrays. Thin Solid Films 2009, 517, 6050−6053. (38) Sun, Z.; Li, Y.; Zhang, J.; Li, Y.; Zhao, Z.; Zhang, K.; Zhang, G.; Guo, J.; Yang, B. A universal approach to fabricate various nanoring arrays based on a colloidal-crystal-assisted-lithography strategy. Adv. Funct. Mater. 2008, 18, 4036−4042. (39) Winzer, M.; Kleiber, M.; Dix, N.; Wiesendanger, R. Fabrication of nano-dot- and nano-ring-arrays by nanosphere lithography. Appl. Phys. A: Mater. Sci. Process. 1996, 63, 617−619.

REFERENCES

(1) Ulman, A. Formation and structure of self-assembled monolayers. Chem. Rev. 1996, 96, 1533−1554. (2) Schreiber, F. Structure and growth of self-assembling monolayers. Prog. Surf. Sci. 2000, 65, 151−256. (3) Poirier, G. E.; Pylant, E. D. The self-assembly mechanism of alkanethiols on Au(111). Science 1996, 272, 1145−1148. (4) Sagiv, J. Organized monolayers by adsorption. 1. Formation and structure of oleophobic mixed monolayers on solid surfaces. J. Am. Chem. Soc. 1980, 102, 92−98. (5) Wen, K.; Maoz, R.; Cohen, H.; Sagiv, J.; Gibaud, A.; Desert, A.; Ocko, B. M. Postassembly chemical modification of a highly ordered organosilane multilayer: New insights into the structure, bonding, and dynamics of self-assembling silane monolayers. ACS Nano 2008, 2, 579−599. (6) Brandow, S. L.; Chen, M.-S.; Aggarwal, R.; Dulcey, C. S.; Calvert, J. M.; Dressick, W. J. Fabrication of patterned amine reactivity templates using 4-chloromethylphenylsiloxane self-assembled monolayer films. Langmuir 1999, 15, 5429−5432. (7) Kim, S.-J.; Ryu, K.; Chang, S. W. Solution-processed organic field-effect transistors patterned by self-assembled monolayers of octadecyltrichlorosilane and phenyltrichlorosilane. J. Mater. Sci. 2010, 45, 566−569. (8) Dulcey, C. S.; Georger, J. H.; Krauthamer, V.; Stenger, D. A.; Fare, T. L.; Calvert, J. M. Deep UV photochemistry of chemisorbed monolayers - patterned coplanar molecular assemblies. Science 1991, 252, 551−554. (9) Sun, S. Q.; Montague, M.; Critchley, K.; Chen, M.-S.; Dressick, W. J.; Evans, S. D.; Leggett, G. J. Fabrication of biological nanostructures by scanning near-field photolithography of chloromethylphenylsiloxane monolayers. Nano Lett. 2006, 6, 29−33. (10) Brandow, S. L.; Schull, T. L.; Martin, B. D.; Guerin, D. C.; Dressick, W. J. Use of low-temperature thermal alkylation to eliminate ink migration in microcontact printed patterns. Chem.-Eur. J. 2002, 8, 5363−5367. (11) Perkins, F. K.; Dobisz, E. A.; Brandow, S. L.; Calvert, J. M.; Kosakowski, J. E.; Marrian, C. R. K. Fabrication of 15 nm wide trenches in Si by vacuum scanning tunneling microscope lithography of an organosilane self-assembled film and reactive ion etching. Appl. Phys. Lett. 1996, 68, 550−552. (12) Marrian, C. R. K.; Perkins, F. K.; Brandow, S. L.; Koloski, T. S.; Dobisz, E. A.; Calvert, J. M. Low voltage electron beam lithography in self-assembled ultrathin films with the scanning tunneling microscope. Appl. Phys. Lett. 1994, 64, 390−392. (13) Dressick, W. J.; Dulcey, C. S.; Brandow, S. L.; Witschi, H.; Neeley, P. F. Proximity x-ray lithography of siloxane and polymer films containing benzyl chloride functional groups. J. Vac. Sci. Technol., A 1999, 17, 1432−1440. (14) Koloski, T. S.; Dulcey, C. S.; Haralson, Q. J.; Calvert, J. M. Nucleophilic displacement reactions at benzyl halide self-assembled monolayer film surfaces. Langmuir 1994, 10, 3122−3133. (15) Mineo, P.; Motta, A.; Lupo, F.; Renna, L.; Gulino, A. Si(111) surface engineered with ordered nanostructures by an atom transfer radical polymerization. J. Phys. Chem. C 2011, 115, 12293−12298. (16) Kimmerlin, T.; Seebach, D. “100 years of peptide synthesis”: ligation methods for peptide and protein synthesis with applications to beta-peptide assemblies. J. Pept. Res. 2005, 65, 229−260. (17) Lu, C. H.; Zhou, W. H.; Han, B.; Yang, H. H.; Chen, X.; Wang, X. R. Surface-imprinted core-shell nanoparticles for sorbent assays. Anal. Chem. 2007, 79, 5457−5461. (18) Facchetti, A.; van der Boom, M. E.; Abbotto, A.; Beverina, L.; Marks, T. J.; Pagani, G. A. Design and preparation of zwitterionic organic thin films: self-assembled siloxane-based, thiophene-spaced Nbenzylpyridinium dicyanomethanides as nonlinear optical materials. Langmuir 2001, 17, 5939−5942. (19) Gu, J. H.; Yam, C. M.; Li, S.; Cai, C. Z. Nanometric protein arrays on protein-resistant monolayers on silicon surfaces. J. Am. Chem. Soc. 2004, 126, 8098−8099. G

dx.doi.org/10.1021/la4010032 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

(40) Abdelsalam, M. E.; Bartlett, P. N.; Baumberg, J. J.; Coyle, S. Preparation of arrays of isolated spherical cavities by self-assembly of polystyrene spheres on self-assembled pre-patterned macroporous films. Adv. Mater. 2004, 16, 90−93. (41) Chen, T.; Chang, D. P.; Jordan, R.; Zauscher, S. Colloidal lithography for fabricating patterned polymer-brush microstructures. Beilstein J. Nanotechnol. 2012, 3, 297−403. (42) Jiang, P.; Hwang, K. S.; Mittleman, D. M.; Bertone, J. F.; Colvin, V. L. Template-directed preparation of macroporous polymers with oriented and crystalline arrays of voids. J. Am. Chem. Soc. 1999, 121, 11630−11637. (43) Lewandowski, B. R.; Kelley, A. T.; Singleton, R.; Lowry, M.; Warner, I. M.; Garno, J. C. Nanostructures of cysteine-coated CdS nanoparticles produced with “two-particle” lithography. J. Phys. Chem. C 2009, 113, 5933−5940. (44) Glangchai, L. C.; Caldorera-Moore, M.; Shi, L.; Roy, K. Nanoimprint lithography based fabrication of shape-specific, enzymatically-triggered smart nanoparticles. J. Controlled Release 2008, 125, 263−272. (45) Chen, J. X.; Liao, W. S.; Chen, X.; Yang, T. L.; Wark, S. E.; Son, D. H.; Batteas, J. D.; Cremer, P. S. Evaporation-induced assembly of quantum dots into nanorings. ACS Nano 2009, 3, 173−180. (46) Kuo, C. W.; Shiu, J. Y.; Chen, P. L.; Somorjai, G. A. Fabrication of size-tunable large-area periodic silicon nanopillar arrays with sub-10nm resolution. J. Phys. Chem. B 2003, 107, 9950−9953. (47) Holland, B. T.; Blanford, C. F.; Do, T.; Stein, A. Synthesis of highly ordered, three-dimensional, macroporous structures of amorphous or crystalline inorganic oxides, phosphates, and hybrid composites. Chem. Mater. 1999, 11, 795−805. (48) McLellan, J. M.; Geissler, M.; Xia, Y. N. Edge spreading lithography and its application to the fabrication of mesoscopic gold and silver rings. J. Am. Chem. Soc. 2004, 126, 10830−10831. (49) Geissler, M.; McLellan, J. M.; Chen, J.; Xia, Y. N. Side-by-side patterning of multiple alkanethiolate monolayers on gold by edgespreading lithography. Angew. Chem., Int. Ed. 2005, 44, 3596−3600. (50) Geissler, M.; McLellan, J. M.; Xia, Y. N. Edge-spreading lithography: Use of patterned photoresist structures to direct the spreading of alkanethiols on gold. Nano Lett. 2005, 5, 31−36. (51) Li, J.-R.; Lusker, K. L.; Yu, J. J.; Garno, J. C. Engineering the spatial selectivity of surfaces at the nanoscale by patterning organosilane self-assembled monolayers via particle lithography. ACS Nano 2009, 3, 2023−2035. (52) Lusker, K. L.; Yu, J.-J.; Garno, J. C. Particle lithography with vapor deposition of organosilanes: A molecular toolkit for studying confined surface reactions in nanoscale liquid volumes. Thin Solid Films 2011, 7, 5223−5229. (53) Saner, C. K.; Lusker, K. L.; LeJeune, Z. M.; Serem, W. K.; Garno, J. C. Self-assembly of octadecyltrichlorosilane: Surface structures formed using different protocols of particle lithography. Beilstein J. Nanotechnol. 2012, 3, 114−122. (54) Nečas, D.; Klapetek, P. Gwyddion: an open-source software for SPM data analysis (http://gwyddion.net/). Cent. Eur. J. Phys. 2012, 10, 181−188. (55) Li, J. R.; Garno, J. C. Elucidating the role of surface hydrolysis in preparing organosilane nanostructures via particle lithography. Nano Lett. 2008, 8, 1916−1922. (56) Liu, G. Y.; Xu, S.; Qian, Y. L. Nanofabrication of self-assembled monolayers using scanning probe lithography. Acc. Chem. Res. 2000, 33, 457−466. (57) Rozlosnik, N.; Gerstenberg, M. C.; Larsen, N. B. Effect of solvents and concentration on the formation of a self-assembled monolayer of octadecylsiloxane on silicon (001). Langmuir 2003, 19, 1182−1188. (58) Vallant, T.; Brunner, H.; Mayer, U.; Hoffmann, H.; Leitner, T.; Resch, R.; Friedbacher, G. Formation of self-assembled octadecylsiloxane monolayers on mica and silicon surfaces studied by atomic force microscopy and infrared spectroscopy. J. Phys. Chem. B 1998, 102, 7190−7197.

(59) Jeon, N. L.; Finnie, K.; Branshaw, K.; Nuzzo, R. G. Structure and stability of patterned self-assembled films of octadecyltrichlorosilane formed by contact printing. Langmuir 1997, 13, 3382−3391. (60) Brandow, S. L.; Chen, M.-S.; Dulcey, C. S.; Dressick, W. J. Formation of aromatic siloxane self-assembled monolayers. Langmuir 2008, 24, 3888−3896. (61) Chen, M.-S.; Brandow, S. L.; Schull, T. L.; Chrisey, D. B.; Dressick, W. J. A non-covalent approach for depositing spatially selective materials on surfaces. Adv. Funct. Mater. 2005, 15, 1364− 1375.

H

dx.doi.org/10.1021/la4010032 | Langmuir XXXX, XXX, XXX−XXX